Abstract:
A series of results similar to Marcinkiewicz's theorem on the interpolation of operators is put forward. The difference from the classical forms of this theorem is that spaces of integrable functions are replaced by some function classes that are extensions of various Hardy spaces.
Some applications of these results to the extension of Carleson's embedding theorem and the Hardy–Littlewood inequalities for analytic functions in Hardy classes are presented.
Bibliography: 41 titles.
Let $(X,\mu)$ be a set with $\sigma$-finite measure $\mu$ and $L^0(X)$ be the set (of equivalence classes) of measurable complex-valued functions on $X$.
We denote by $L^{p_0}(X)+L^{p_1}(X)$ the space of functions $f\in L^0(X)$ representable in the form $f=f_0+f_1$, where $f_0\in L^{p_0}(X)$ and $f_1\in L^{p_1}(X)$. We also set
Let $(Y,\nu)$ be another set with $\sigma$-finite measure $\nu$, and let $T$ be an operator with values in $L^0(Y)$ defined on some subset of $L^0(X)$.
The operator $T$ is called quasi-subadditive if there exists a positive number ${K_1=K_1(T)}$ such that1[x]1Here and in what follows all inequalities for functions are pointwise.
then $T$ is quasilinear ($T(u+v)$ is assumed to be defined whenever $Tu$ and $Tv$ are; $T(\lambda u)$ is defined whenever $Tu$ is). In the case when $K_1=1$ we use the terms ‘subadditive’ and ‘sublinear’, respectively.
We write $A\lesssim B$ if $A\leqslant CB$ for some positive constant $C$, which can depend on some parameters, which will be indicated explicitly.
1.2. Marcinkiewicz’s theorem
In the original version of Marcinkiewicz’s interpolation theorem for $L^p$-spaces, the linear or positive subadditive operator $T$ was assumed to satisfy a weak-type condition of the form
Theorem 1. Let $T\colon L^{p_0}(X)+L^{p_1}(X)\to L^0(X)$ be a quasi-subadditive operator, and let $1\leqslant p_0\leqslant q_0\leqslant\infty$, $1\leqslant p_1\leqslant q_1\leqslant\infty$, $p_0<p_1$ and $q_0\ne q_1$.
Assume that there exist constants $M_0$ and $M_1$ such that
($\lesssim$ depends on $K_1$, $p_0$, $p_1$, $q_0$, $q_1$ and $\theta$).
This result is commonly referred to as Marcinkiewicz’s interpolation theorem; this is an important tool in many branches of modern analysis involving $L^p$-spaces of summable functions.
The first step towards Theorem 1 was the note [1] by Marcinkiewicz, which puts forward, without proof, its ‘diagonal’ variant $p_0=q_0$, $p_1=q_1$. The complete version of the result was later established by Zygmund [2].
1.3. The Stein–Weiss–Grafakos theorem
An application of condition (1.5) to the characteristic function $\chi_A$ of a measurable set $A\subset X$ produces the inequality
which is called a (restricted) weak-type inequality.
In Marcinkiewicz’s interpolation theorem such conditions were originally introduced by Stein and Weiss [7] (see also § 5.3 in [8]), who assumed in [7] that $p_0,p_1,q_0,q_1\geqslant 1$ and considered the operators defined on the class $S(X)$ of simple functions on $X$ (finite linear combinations of characteristic functions of measurable sets of finite measure).
Grafakos (see [9], Theorem 1.4.19) extended Marcinkiewicz’s interpolation theorem with the Stein–Weiss conditions (1.8) to arbitrary positive parameters $p$ and $q$. To formulate his result we recall the definition of Lorentz spaces.
For $0<p,r\leqslant\infty$ we let $L^{p,r}(X)$ denote the Lorentz space (see [10] and also [8], § 5.3) with the quasinorm
(for example, see [8], § 5.3). For $r=\infty$ the quasinorm (1.9) coincides with (1.6) (see [8], Lemma 3.8).
Theorem 2. Let $0<p_0\ne p_1\leqslant\infty$, $0<q_0\ne q_1\leqslant\infty$ and $0<r<\infty$, and let $T\colon L^{p_0}(X)+L^{p_1}(X)\to L^0(Y)$ be a quasilinear continuous operator satisfying the following condition: there exist constants $M_0$ and $M_1$ such that, for each measurable set $A\subset X$ of finite measure
($\lesssim$ depends on $K_1$, $p_0$, $p_1$, $q_0$, $q_1$, $r$ and $\theta$).
Theorem 1.4.19 in [9] is close to Theorem 2, but there the operator $T$ is not assumed to be continuous. However, the proof in [9] contained a gap, which was filled in [11], where both the results and proofs were corrected. The validity of Theorem 2 was pointed out in [11] (see Remark 1.4, (iii), (iv)), where its variant for operators defined on some subclass of the class $S(X)$ of simple functions on $X$ was also proved (Theorem 1.1 in [11]). The proof there is similar to the proof of Theorem 1.4.19 in [9]. A related version of the same theorem was later established by Grafakos [12] (see Theorem 1.4.19 there). The condition of local continuity in related problems (a multilinear version of Marcinkiewicz’s theorem) was used in [13].
Our aim here is to obtain analogues of Theorem 2 for operators satisfying weak-type $L^p$-inequalities and acting on some spaces which are natural extensions of the Hardy classes $H^p$ of analytic or harmonic functions. Some applications of these results will also be given.
§ 2. The main results
2.1. The spaces $\mathcal H^p(\mathbf X)$
Let $X$ be a Hausdorff space whose topology is induced by a quasimetric $d$, that is, $d\colon X\times X\to[0,\infty)$ is a function satisfying all axioms of a metric except the triangle inequality, which is replaced by a weaker condition: there exists a number $K_2=K_2(d)\geqslant 1$ such that, for all $x,y,z\in X$,
for each function $u\colon \mathbf X\to\mathbb{C}$.
We let $\mathcal H^0(\mathbf X)$ denote the set of all measurable functions $u\colon \mathbf X\to\mathbb{C}$ whose maximal function $\mathcal N u$ is finite $\mu$-almost everywhere (equivalent functions are not identified). We equip $\mathcal H^0(\mathbf X)$ with the metric
For $r=p$ we write $\mathcal H^p(\mathbf X)$ in place of $\mathcal H^{p,p}(\mathbf X)$.
In the case $\mathbf X=\mathbb R^{n+1}_+$ the classes $\mathcal H^p(\mathbb R^{n+1}_+)$ were considered for the first time in [15], where it was assumed in addition that functions in these classes are continuous and have nontangential limits almost everywhere (see also [16] and [17], where the case of general $\mathbf X$ was considered).
The concept of a maximal function and the condition $\mathcal N u\in L^p(X)$, which date back to Hardy and Littlewood [18], are widely useful in the theory of Hardy spaces (see, for example, [19] for $\mathbf X=\mathbb R_+^{n+1}$, and Theorem 5.6.5 in [20], where $\mathbf X$ is the unit ball in $\mathbb{C}^n$). A number of boundary-value problems are formulated in terms of maximal functions (see, for example, [21] and [22]; this list is far from complete). This comment describes the possible range of application of our results.
2.2. The main results
By $\mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)$ we denote the space of all functions $u\in \mathcal H^0(\mathbf X)$ representable as $u=u_0+u_1$, where $u_0\in \mathcal H^{p_0}(\mathbf X)$ and $u_1\in \mathcal H^{p_1}(\mathbf X)$. We equip this space with the quasinorm
is used in our main results, which are as follows.
Theorem 3. Let $0<p_0\ne p_1\leqslant\infty$, $0<q_0\ne q_1\leqslant\infty$ and $0<r<\infty$, and let $T$ be a quasilinear continuous operator from $\mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)$ to $L^0(Y)$ which satisfies the following conditions: there exist positive constants $M_0$ and $M_1$ such that, for each measurable set $A\subset\mathbf X$ satisfying $\mu(\widehat{A})<\infty$,
($\lesssim$ depends on $K_1$, $p_0$, $p_1$, $q_0$, $q_1$, $r$ and $\theta$).
Theorem 3 becomes Theorem 2 if $\mathbf X$, $\mathcal H^{p}(\mathbf X)$ and $\widehat{A}$ are replaced by $X$, $L^p(X)$ and $A\subset X$, respectively.
Corollary 1. Under the assumptions of Theorem 3, let $\theta\in(0,1)$, and let $p$ and $q$, $p\leqslant q$, be defined by (1.7).
Then for each function $u\in\mathcal H^{p}(\mathbf X)$,
($\lesssim$ depends on $K_1$, $p_0$, $p_1$, $q_0$, $q_1$, $\theta$).
Corollary 2 is a consequence of Theorem 3 with $r=p$.
The next theorem is an analogue of the classical form of Marcinkiewicz’s theorem (see Theorem 1).
Theorem 4. Let $0<p_0\ne p_1\leqslant\infty$, $0<q_0\ne q_1\leqslant\infty$, and let the quasilinear operator $T\colon \mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)\to L^ 0(Y)$ satisfy the following condition: there exist positive constants $M_0$ and $M_1$ such that, for all $\lambda>0$,
($\lesssim$ depends on $K_1$, $p_0$, $p_1$, $q_0$, $q_1$, $\theta$).
Now we state an analogue of Theorem 3 for operators acting from ${\mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)}$ to $\mathcal H^0(\mathbf Y)$. Here $\mathbf Y=Y\times(0,\tau_0)$, where $(Y,\nu)$ is as above and $0<\tau_0\leqslant\infty$. In addition, $Y$ is equipped with a quasimetric $d_Y$ generating the maximal function
which, in its turn, generates the space $\mathcal H^0(\mathbf Y)$.
Theorem 5. Let $0<p_0\ne p_1\leqslant\infty$, $0<q_0\ne q_1\leqslant\infty$ and $0<r<\infty$, and let $T$ be a quasilinear continuous operator from $\mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)$ to $\mathcal H^0(\mathbf Y)$ satisfying the following condition: there exist positive constants $M_0$ and $M_1$ such that for each measurable set $A\subset\mathbf X$ such that $\mu(\widehat{A})<\infty$,
We need the following properties of Lorentz spaces, which were defined in § 1.3 above.
The scale of Lorentz spaces is monotone in the second parameter (see, for example, [9], Proposition 1.4.10, or [8], Theorem 3.11): if $0<p\leqslant\infty$ and $0< q_0<q_1\leqslant\infty$, then $L^{p,q_0}(X)\subset L^{p,q_1}(X)$ and
to power $p$ and averaging over $y\in B(x,t)$ we obtain
$$
\begin{equation}
|u(x,t)|\leqslant\biggl(\frac{1}{\mu(B(x,t))}\int_{B(x,t)}(\mathcal N u)^p\,d\mu\biggr)^{1/p}\lesssim [\mu(B(x,t))]^{-1/p}\|\mathcal N u\|_{L^p(X)}.
\end{equation}
\tag{3.7}
$$
Now let $\{u_n\}$ be a Cauchy sequence in $\mathcal H^p(\mathbf X)$. By (3.7), $\{u_n\}$ converges at each point $(x,t)\in\mathbf X$ to some measurable function $u$ on $\mathbf X$.
Let $\{n_k\}$ be an increasing sequence of indices such that
Hence, by the choice of $\{n_k\}$ the sequence $\{u_{n_k}\}$ converges to $u$ in $\mathcal H^p(\mathbf X)$, and therefore the whole sequence $\{u_n\}$ has the same property. Lemma 2 is proved.
In what follows, we frequently use the following notation in proofs:
$$
\begin{equation}
E[\lambda]=\{x\in X\colon \mathcal N u(x)>\lambda\},\qquad\lambda>0,
\end{equation}
\tag{3.8}
$$
Proof. (1) If $x\in\widehat{\mathcal T(E)}$, then $D(x)\cap\mathcal T(E)\ne\varnothing$, but if $x\notin E$, then $D(x)\cap\mathcal T(E)=\varnothing$ since
Hence $\widehat{\mathcal T(E)}\subset E$. Conversely, if $x\in E$, but $x\notin\widehat{\mathcal T(E)}$, then $D(x)\cap \mathcal T(E)=\varnothing$, that is, $D(x)\subset [\mathcal T(E)]^{\mathrm c}$.
is always present in the proofs of various variants of Marcinkiewicz’s interpolation theorem.
The following lemma, albeit simple, is important because it shows how the nontangential maximal function $\mathcal N u$ is transformed with the splitting (3.10).
Lemma 4. If $u\colon \mathbf X\to\mathbb{C}$ and $\lambda>0$, then
$$
\begin{equation*}
\mathcal N u^\lambda(x) = \begin{cases} \mathcal N u(x) & \textit{if } \mathcal N u(x)>\lambda, \\ 0 & \textit{if }\mathcal N u(x)\,{\leqslant}\,\lambda, \end{cases}
\end{equation*}
\notag
$$
and
$$
\begin{equation*}
\mathcal N u_\lambda(x) \leqslant \begin{cases} \lambda & \textit{if } \mathcal N u(x)>\lambda, \\ \mathcal N u(x) &\textit{if }\mathcal N u(x)\leqslant\lambda. \end{cases}
\end{equation*}
\notag
$$
Proof. If $\mathcal N u(x)>\lambda$, then $\mathcal N u(x):=\mathcal N u^\lambda(x)$, because the supremum in the definitions of $\mathcal N u(x)$ and $\mathcal N u^\lambda(x)$ is taken over the same set
and $\mathcal N u_\lambda(x)\leqslant\lambda$, because $|u_\lambda(y,t)|\leqslant\lambda$ for all $(y,t)\in\mathbf X$.
If $\mathcal N u(x)\leqslant\lambda$, then $|u(y,t)|\leqslant \lambda$ for $(y,t)\in D(x)$, and for these $(y,t)$, we have $u^\lambda(y,t)=0$ and $u_\lambda(y,t)=u(y,t)$.
This proves the lemma.
3.5. Kalton’s lemma
The following lemma, which is a key ingredient in the proof of Theorem 3, is similar to Lemma 1.4.20 in [9], which (as noted in [9], p. 74) was proposed by Kalton.
Lemma 5. Let $0<p<\infty$ and $0<q\leqslant\infty$, and let $T$ be a quasilinear continuous operator $\mathcal H^{p}(\mathbf X)\to L^0(Y)$ satisfying the following condition: there exists a positive constant $M$ such that
($\lesssim$ depends on $K_1$, $p$, $q$, $\alpha$).
Proof. First we note that if $\alpha_1$ is a solution of the equation $(2K_1)^{\alpha_1}=2$, then by Lemma 1, for all $0<\alpha\leqslant\alpha_1$ and $\{v_k\}_{k=1}^{m}\subset \mathcal H^p(\mathbf X)$, we have the pointwise inequality
Now let $0<\alpha_0\leqslant\alpha_1$ be a number such that $\alpha_0<q$. Then $q/\alpha>1$ for each $0<\alpha\leqslant\alpha_0$, and so the quasinormed Lorentz space $L^{q/\alpha,\infty}(Y)$ is normed (see § 3.2): the norm
Let us now prove the following key fact: there exists a constant $C=C(q,\alpha)$ such that for each nonnegative bounded function $u\in\mathcal H^p(\mathbf X)$ and any measurable set $A\subset\mathbf X$ satisfying $\mu(\widehat{A})<\infty$,
is the sequence of partial sums of series (3.16), then by the continuity of the operator $T$ the sequence $\{T(u\chi_A-u_m)\}$ converges to zero in measure on $A$, and some subsequence $\{T(u\chi_A- u_{m_k})\}$ of it converges to zero almost everywhere.
Since the operator $T$ is quasilinear (see (1.2) and (1.3)), it follows from (3.14) that
Let us now verify (3.13) for $u\in\mathcal H^{p}(\mathbf X)$. Using the notation (3.8) and (3.9), consider the sets
$$
\begin{equation*}
\Delta_n=\mathcal T(E[(\mathcal N u)^*(2^{n})])\setminus\mathcal T(E[(\mathcal N u)^*(2^{n+1})]),\qquad n\in\mathbb Z.
\end{equation*}
\notag
$$
$$
\begin{equation*}
\mathcal N(u\chi_{(\mathbf X\setminus\mathcal T_{-M})})=\mathcal N u \quad\text{on the set}\ \{\mathcal N u\leqslant2^{-M}\}
\end{equation*}
\notag
$$
and
$$
\begin{equation*}
\mathcal N(u\chi_{(\mathbf X\setminus\mathcal T_{-M})})\leqslant 2^{-M} \quad\text{on the set } \{\mathcal N u>2^{-M}\}.
\end{equation*}
\notag
$$
Hence
$$
\begin{equation*}
\begin{aligned} \, & \|u\chi_{(\mathbf X\setminus\mathcal T_{-M})}\|_{\mathcal H^p(\mathbf X)}^p=\int_{\{\mathcal N u\leqslant2^{-M}\}}(\mathcal N u)^p\,d\mu+\int_{\{\mathcal N u>2^{-M}\}}(\mathcal N u)^p\,d\mu \\ &\qquad \leqslant\int_{\{\mathcal N u\leqslant2^{-M}\}}(\mathcal N u)^p\,d\mu+2^{-Mp}\mu(\{\mathcal N u>2^{-M}\})\to0,\qquad M\to\infty. \end{aligned}
\end{equation*}
\notag
$$
Consequently, the $\mathcal H^p(\mathbf X)$-quasinorms of both terms on the right in (3.22) converge to zero as $N,M\to\infty$, which proves the $\mathcal H^p(\mathbf X)$-convergence of the series (3.21).
The operator $T$ is continuous, and so, proceeding as in the proof of (3.17), we get that
To estimate the last sum it suffices to invoke (3.6). Thus, we have proved the conclusion of Lemma 5 for nonnegative functions $u\in\mathcal H^p(\mathbf X)$.
We can represent any complex-valued function $u$ as $u=u_1-u_2+i(v_1-v_2)$, where the nonnegative functions $u_j$ and $v_j$, $j=1,2$, are defined by
We proceed as in the proof of Theorem 2 (see [9], § 1.4.4). Let $0<p_0<p_1<\infty$ (the case $p_1=\infty$ is considered separately). First, using conditions (1.10) and (1.11) and Lemma 5, we find that
and for any $t>0$ we represent $u\in\mathcal H^{p,r}(\mathbf X)$ as the sum $u=u^\lambda+u_\lambda$ (see (3.10)), where $\lambda:=(\mathcal N u)^*(\delta t^{\gamma})$ and $\delta>0$ is a number to be chosen below from some optimization considerations.
From (3.10), Lemma 4, and the definition of an equimeasurable rearrangement we obtain
$$
\begin{equation}
\begin{cases} (\mathcal N u^\lambda)^*(s)\leqslant(\mathcal N u)^*(s) &\text{for }0<s\leqslant\delta t^\gamma, \\ (\mathcal N u^\lambda)^*(s)=0 & \text{for } s>\delta t^\gamma, \\ (\mathcal N u_\lambda)^*(s)=(\mathcal N u)^*(\delta t^\gamma) & \text{for } 0<s\leqslant\delta t^\gamma, \\ (\mathcal N u_\lambda)^*(s)\leqslant (\mathcal N u)^*(s) & \text{for } s>\delta t^\gamma. \end{cases}
\end{equation}
\tag{4.3}
$$
In particular, by (4.3) we have $u^\lambda\in\mathcal H^{p_0,\beta}(\mathbf X)$ and $u_\lambda\in\mathcal H^{p_1,\beta}(\mathbf X)$ for each $t>0$.
Let us now estimate $\|Tu\|_{L^{q,r}(Y)}$. Since the operator $T$ is quasilinear, we have
$$
\begin{equation*}
\begin{aligned} \, \biggl\|t^{1/q}(Tu^\lambda)^*\biggl(\frac t2\biggr)\biggr\|_{L^r(dt/t)} &=\biggl\|t^{1/q-1/q_0}t^{1/q_0}(Tu^\lambda)^*\biggl(\frac t2\biggr)\biggr\|_{L^r(dt/t)} \\ &\leqslant2^{1/q_0}C_0M_0\|t^{1/q-1/q_0}\|\mathcal N u^\lambda\|_{L^{p_0,\beta}(X)}\|_{L^r(dt/t)} \\ &=2^{1/q_0}C_0M_0\|t^{\gamma(1/p_0-1/p)}\|\mathcal N u^\lambda\|_{L^{p_0,\beta}(X)}\|_{L^r(dt/t)}. \end{aligned}
\end{equation*}
\notag
$$
Changing the variable to $\tau:=\delta t^\gamma$, employing the first two relations in (4.3) and applying Hardy’s inequality (3.1) for $p=r/\beta\geqslant 1$ and $b=r(1/p_0-1/p)$, we continue the estimate as follows:
Substituting estimates (4.7) and (4.8) into the right-hand side of (4.4) we arrive at the inequality
$$
\begin{equation*}
\|Tu\|_{L^{q,r}(Y)}\leqslant C\bigl(M_0\delta^{1/p_0-1/p}+M_1\delta^{1/p-1/p_1}\bigr) \|\mathcal N u\|_{L^{p,r}(X)},
\end{equation*}
\notag
$$
where $C=C(K_1,p_0,p_1,q_0,q_1,r,\theta)$ is a positive constant depending only on the above parameters. Now we choose $\delta>0$ so as to minimize the right-hand side of the last inequality. This is equivalent to saying that
This proves inequality (2.6) in the case $0<p_1<\infty$.
The case $p_1=\infty$ can be reduced to the one just considered as follows. We set $p^*\in(p,\infty)$ and choose $\theta^*\in(0,1)$ so as to have
$$
\begin{equation*}
\frac{1}{p^*}=\frac{1-\theta^*}{p_0}+\frac{\theta^*}{\infty}; \quad\text{we also set}\ \frac{1}{q^*}:=\frac{1-\theta^*}{q_0}+\frac{\theta^*}{q_1}.
\end{equation*}
\notag
$$
Let $g\in L^0(Y)$, and let $\nu_{g}(\lambda):=\nu\{x\in X\colon |g(x)|>\lambda\}$ be the distribution function of $g$. The following inequality is clear:
This means that the operator $T$ is of restricted weak type $(p^*,q^*)$. An application of what has been proved for the pairs $(p_0,q_0)$ and $(p^*,q^*)$ shows that
This result is a direct consequence of Theorem 3, since by the substitution of the characteristic functions of sets into inequalities (2.8) and (2.9) we obtain (2.4) and (2.5), respectively. The continuity of the operator $T\colon \mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)\to L^0(Y)$ is secured by (2.8) and (2.9).
5.1. The Carleson–Duren–Hörmander embedding theorem
A Borel measure $\nu$ on $\mathbf X$ is said to be a Carleson measure of order $\alpha\!>\!0$ (written ${\nu\mkern-1mu\!\in\!\mkern-1mu CM_{\alpha}\mkern-1mu(\mathbf X)}$) if
Lemma 6. Let the measure $\mu$ on $X$ satisfy the doubling condition (5.2), and let $\nu\in CM_{\alpha}(\mathbf X)$ for some $\alpha\geqslant1$. Then for each measurable set $A\subset\mathbf X$ such that $\mu(\widehat{A})<\infty$,
If $x\in\widehat{A_n}$ (recall that $\widehat{A_n}$ is defined in (2.3)), then there exists a point $(y,t)\in\mathbf X$ such that
$$
\begin{equation*}
(y,t)\in D(x)\cap A, \qquad 0<t\leqslant n
\end{equation*}
\notag
$$
(see (2.2)). We set $B_x:=B(y,t)\subset\widehat{A_n}$. The family of balls $\{B_x\colon x\in\widehat{A_n}\}$ of uniformly bounded radii covers $\widehat{A_n}$, and this family contains a finite or countable subfamily $\{B_k:=B_{x_k}\}$ such that
In the case when $t_0<\infty$ the arguments are simpler, since there is no need to introduce the sequence $\{A_n\}$. Lemma 6 is proved.
Theorem 6. Let $0<p\leqslant q<\infty$, let the measure $\mu$ on $X$ satisfy the doubling condition, and let $\nu$ be a measure on $\mathbf X$ whose domain contains that of the measure $\mu\times m_1$. Then the following conditions are equivalent:
(1) $\nu$ is a Carleson measure of order $q/p$ on $\mathbf X$;
(2) inequality (5.3) holds for each measurable set $A\subset\mathbf X$, $\mu(\widehat{A})<\infty$ ($\lesssim$ does not depend on $A$);
(3) for all $u\in\mathcal H^p(\mathbf X)$ and $\lambda>0$,
Consider the identity operator $\mathrm{Id}\colon \mathcal H^{p_0}(\mathbf X)+\mathcal H^{p_1}(\mathbf X)\to L^0(\mathbf X)$. Clearly, this operator is linear and continuous. We apply Lemma 6 to each of the pairs $(p_0,q_0)$ and $(p_1,q_1)$ (for these pairs (5.3) clearly corresponds to conditions (2.4) and (2.5) for this operator). Now Theorem 6 follows from Corollary 1.
Let us illustrate Theorem 6 with two particular cases.
Example 1. For $n\geqslant 1$ let $B^n\subset\mathbb{C}^n$ be the open unit ball in $\mathbb{C}^n$. Let $X=S=\partial B^n\subset\mathbb{C}^n$ be the unit sphere (the boundary of $B^n$) and $\mu=\sigma$ be the surface Lebesgue measure on $S$ normalized by the condition $\sigma(S)=1$. Consider the natural quasimetric
we find that $H^p(B^n)$ for $p>0$ is continuously embedded in $\mathcal H^p(S\times(0,1))$ (see, for example, [20], Theorem 5.6.5). Now Theorem 6 implies that if $0<p\leqslant q<\infty$ and $\nu$ is a Carleson measure on $B^n$ of order $q/p$, then
The first inequality here holds for any measurable function $f\colon S\times(0,1)\to\mathbb{C}$, and the second holds for $f\in H^p(B^n)$.
The definition of a Carleson measure dates back to Carleson’s papers [30] and [31], where inequalities of the form (5.7) (for $n=1$ and $q=p$) were used to solve important problems on Hardy classes of analytic functions on the unit disc (namely, the problem of interpolation by analytic functions and the corona problem). For more details, see also [32], Chs. 7 and 8.
Hörmander [33] extended inequality (5.7) for $q=p$ to the Hardy classes of holomorphic functions on domains in $\mathbb{C}^n$, $n\geqslant1$, with sufficiently smooth boundaries; the diagonal variant of Marcinkiewicz’s interpolation theorem was used in the proof.
Duren [34] employed Hörmander’s method in the one-dimensional case with ${0<p\leqslant q<\infty}$ (also see the discussion of [33] and [34] in [28], § 1.4).
Example 2. Let $X=\mathbb R^n$, $n\geqslant 1$, let $d(x,y)=|x-y|$ be the Euclidean metric, and $\mu$ be the Lebesgue measure on $\mathbb R^n$, $I=(0,\infty)$.
Here, one can take as the Hardy class $H^p(\mathbb R^{n+1}_+)$, $p>0$, for example, the set of harmonic functions $u$ in $\mathbb R^{n+1}_+:=\mathbb R^n\times\mathbb R_+$ whose nontangential maximal function $\mathcal N u$ lies in $L^p(\mathbb R^n)$.
The following more general variant of the classes $H^p$ can also be used. Let $\varphi$ be a sufficiently smooth function on $\mathbb R^n$ (for example, $\varphi$ lies in the Schwartz class), and let
Let $H^p_{\varphi}(\mathbb R^{n+1}_+)$ be the class of convolutions $u=f\ast \varphi_t$ of tempered distributions $f$ whose nontangential maximal function $\mathcal N(f\ast \varphi_t)$ lies in $L^p(\mathbb R^n)$ (for more details, see [19] or [35], Ch. 2). Then
The possible use of nontangential maximal functions in the proof of Carleson’s embedding theorem for Poisson integrals of $L^p(\mathbb R^n)$-functions, $p>1$, was pointed out in [4], Ch. 7, § 4.4.
In this regard we also mention the paper [36], where this idea was used to study Carleson measures on products of the form $X\times(0,\infty)$, where $X$ is a quasimetric space equipped with a measure satisfying the doubling condition (5.2). However, in [36] the functions were additionally assumed to satisfy a Harnack-type inequality.
5.2. Hardy–Littlewood inequalities
Here we consider some inequalities for $\mathcal H^p(\mathbf X)$-functions, $p>0$, which originate from the Hardy–Littlewood inequalities for analytic functions in Hardy classes in the unit disc in $\mathbb{C}$. We set
We illustrate Theorem 7 by using the spaces $X$ from Examples 1 and 2.
For the unit disc $B^1\subset\mathbb{C}$ (see Example 1 with $n=1$) inequalities in Theorem 7 were studied by Hardy and Littlewood (see [37], Theorem 2, [38], Theorems 27 and 31, and [39], Theorems 1 and 11), who proved (5.9)–(5.11) for analytic functions in the Hardy classes $H^p(B^1)$.
Flett (see [40], Theorem 1) established inequalities (5.9)–(5.11) for $p\geqslant1$ for Poisson integral-type operators on some locally compact groups, and applied them, in particular, to Poisson integrals in the half-space $\mathbb R^{n+1}$ for $p\geqslant 1$ (see [40], Theorem 2) and also, for $p>0$ (see [40], Theorem 3), to functions $u$ such that some power $u^k$, $k\leqslant p$, is subharmonic. Flett’s arguments depend on the diagonal version of Marcinkiewicz’s interpolation theorem. His idea was subsequently used by Mitchell and Hahn (see [41], Theorem 4), who carried over the one-dimensional Hardy–Littlewood inequalities to the multivariate case of holomorphic functions in the Hardy classes $H^p$ on the unit ball or on a bounded symmetric domain in $\mathbb{C}^n$. All these results are corollaries to Theorem 7.
Bibliography
1.
J. Marcinkiewicz, “Sur l'interpolation d'opérations”, C. R. Acad. Sci. Paris, 208 (1939), 1272–1273
2.
A. Zygmund, “On a theorem of Marcinkiewicz concerning interpolation of operations”, J. Math. Pures Appl. (9), 35 (1956), 223–248
3.
A. Zygmund, Trigonometric series, v. II, 2nd ed., Cambridge Univ. Press, New York, 1959, vii+354 pp.
4.
E. M. Stein, Singular integrals and differentiability properties of functions, Princeton Math. Ser., 30, Princeton Univ. Press, Princeton, NJ, 1970, xiv+290 pp.
5.
S. G. Kreĭn, Ju. I. Petunin and E. M. Semenov, Interpolation of linear operators, Transl. Math. Monogr., 54, Amer. Math. Soc., Providence, RI, 1982, xii+375 pp.
6.
J. Bergh and J. Löfström, Interpolation spaces. An introduction, Grundlehren Math. Wiss., 223, Springer-Verlag, Berlin–New York, 1976, x+207 pp.
7.
E. M. Stein and G. Weiss, “An extension of a theorem of Marcinkiewicz and some of its applications”, J. Math. Mech., 8:2 (1959), 263–284
8.
E. M. Stein and G. Weiss, Introduction to Fourier analysis on Euclidean spaces, Princeton Math. Ser., 32, Princeton Univ. Press, Princeton, NJ, 1971, x+297 pp.
9.
L. Grafakos, Classical Fourier analysis, Grad. Texts in Math., 249, 2nd ed., Springer, New York, 2008, xvi+489 pp.
10.
G. G. Lorentz, “Some new functional spaces”, Ann. of Math. (2), 51:1 (1950), 37–55
11.
Yi Yu Liang, Li Guang Liu and Da Chun Yang, “An off-diagonal Marcinkiewicz interpolation theorem on Lorentz spaces”, Acta Math. Sin. (Engl. Ser.), 27:8 (2011), 1477–1488
12.
L. Grafakos, Classical Fourier analysis, Grad. Texts in Math., 249, 3rd ed., Springer, New York, 2014, xviii+638 pp.
13.
L. Grafakos and N. Kalton, “Some remarks on multilinear maps and interpolation”, Math. Ann., 319:1 (2001), 151–180
14.
V. I. Bogachev, Measure theory, v. I, Springer-Verlag, Berlin, 2007, xviii+500 pp.
15.
R. R. Coifman, Y. Meyer and E. M. Stein, “Some new function spaces and their applications in harmonic analysis”, J. Funct. Anal., 62:2 (1985), 304–335
16.
V. G. Krotov, “On the boundary behavior of functions in spaces of Hardy type”, Math. USSR-Izv., 37:2 (1991), 303–320
17.
V. G. Krotov, “Tent spaces and their applications”, Theory of functions and approximations: Saratov Winter School, v. 1, Publishing House of Saratov University, Saratov, 1992, 90–102 (Russian)
18.
G. H. Hardy and J. E. Littlewood, “A maximal theorem with function-theoretic applications”, Acta Math., 54:1 (1930), 81–116
19.
C. Fefferman and E. M. Stein, “$H^p$ spaces of several variables”, Acta Math., 129:3–4 (1972), 137–193
20.
W. Rudin, Function theory in the unit ball of $\mathbb C^n$, Grundlehren Math. Wiss., 241, Springer-Verlag, New York–Berlin, 1980, xiii+436 pp.
21.
G. Verchota, “The Dirichlet problem for the polyharmonic equation in Lipschitz domains”, Indiana Univ. Math. J., 39:3 (1990), 671–702
22.
J. Pipher and G. C. Verchota, “Dilation invariant estimates and the boundary Gårding inequality for higher order elliptic operators”, Ann. of Math. (2), 142:1 (1995), 1–38
23.
V. G. Krotov, “Marcinkiewicz interpolation theorem for spaces of Hardy type”, Math. Notes, 113:2 (2023), 306–310
24.
V. G. Krotov, “Interpolation of operators in Hardy-type spaces”, Proc. Steklov Inst. Math., 323 (2023), 173–187
25.
R. A. Hunt, “On $L(p,q)$ spaces”, Enseign. Math. (2), 12 (1966), 249–276
26.
T. Aoki, “Locally bounded linear topological spaces”, Proc. Imp. Acad. Tokyo, 18:10 (1942), 588–594
27.
S. Rolewicz, Metric linear spaces, Math. Appl. (East European Ser.), 20, 2nd ed., D. Reidel Publishing Co., Dordrecht; PWN–Polish Sci. Publ., Warsaw, 1985, xii+459 pp.
28.
A. B. Aleksandrov, “Function theory in the ball”, Several complex variables. II. Function theory in classical domains. Complex potential theory, Encycl. Math. Sci., 8, 1994, 107–178
29.
A. P. Calderón, “Inequalities for the maximal function relative to a metric”, Studia Math., 57:3 (1976), 297–306
30.
L. Carleson, “An interpolation problem for bounded analytic functions”, Amer. J. Math., 80:4 (1958), 921–930
31.
L. Carleson, “Interpolations by bounded analytic functions and the corona problem”, Ann. of Math. (2), 76:3 (1962), 547–559
32.
J. B. Garnett, Bounded analytic functions, Pure Appl. Math., 96, Academic Press, Inc., New York–London, 1981, xvi+467 pp.
33.
L. Hörmander, “$L^p$ estimates for (pluri-) subharmonic functions”, Math. Scand., 20 (1967), 65–78
34.
P. L. Duren, “Extension of a theorem of Carleson”, Bull. Amer. Math. Soc., 75:1 (1969), 143–146
35.
L. Grafakos, Modern Fourier analysis, Grad. Texts in Math., 250, 3rd ed., Springer, New York, 2014, xvi+624 pp.
36.
S. C. Gadbois and W. T. Sledd, “Carleson measures on spaces of homogeneous type”, Trans. Amer. Math. Soc., 341:2 (1994), 841–862
37.
G. H. Hardy and J. E. Littlewood, “A convergence criterion for Fourier series”, Math. Z., 28:1 (1928), 612–634
38.
G. H. Hardy and J. E. Littlewood, “Some properties of fractional integrals. II”, Math. Z., 34:1 (1932), 403–439
39.
G. H. Hardy and J. E. Littlewood, “Theorems concerning mean values of analytic or harmonic functions”, Quart. J. Math. Oxford Ser., 12:1 (1941), 221–256
40.
T. M. Flett, “On the rate of growth of mean values of holomorphic and harmonic functions”, Proc. London Math. Soc. (3), 20:4 (1970), 749–768
41.
J. Mitchell and K. T. Hahn, “Representation of linear functionals in $H^p$ spaces over bounded symmetric domains in ${C}^N$”, J. Math. Anal. Appl., 56:2 (1976), 379–396
Citation:
V. G. Krotov, “Marcinkiewicz's interpolation theorem for Hardy-type spaces and its applications”, Sb. Math., 215:8 (2024), 1091–1113